You are on page 1of 9

Synchronizing Separation Flow

Control With Unsteady Wakes in a


M. Bloxham
D. Reimann
Low-Pressure Turbine Cascade
K. Crapo Particle image velocimetry (PIV) measurements were made on a highly loaded low-
pressure turbine blade in a linear cascade. The Pack B blade has a design Zweifel
J. Pluim coefficient of 1.15 and a peak C p at 63% axial chord on the suction surface. Data were
taken at Rec ⫽ 20 K with 3% inlet freestream turbulence and a wake-passing flow coef-
J. P. Bons ficient of 0.8. Without unsteady wakes, a nonreattaching separation bubble exists on the
suction surface of the blade beginning at 68% axial chord. The time-averaged separation
Department of Mechanical Engineering, zone is reduced in size by approximately 35% in the presence of unsteady wakes. Phase-
Brigham Young University, locked hot-wire and PIV measurements were used to document the dynamics of this
Provo, UT 84602 separation zone when subjected to synchronized, unsteady forcing from a spanwise row of
vortex generator jets (VGJs) in addition to the unsteady wakes. The phase difference
between VGJ actuation and the wake passing was optimized. Both steady state C p and
phase-locked velocity measurements confirm that the optimal combination of wakes and
jets yields the smallest separation. 关DOI: 10.1115/1.2952376兴

Introduction tained by Bons et al. 关8兴 compared favorably to the control


achieved with steady VGJs but at a fraction of the mass flow
Low-pressure turbine 共LPT兲 blades have been shown to be sus-
requirements. These results were obtained over a range of forcing
ceptible to boundary layer separation at low Reynolds numbers
frequencies and duty cycles with the conclusion that both vari-
关1–3兴. Many techniques have been developed to decrease the ex-
ables had little impact on the time-averaged wake losses. The
tent of the separation in an attempt to reduce the total pressure
forcing frequency independence was demonstrated over a forcing
losses. Of these techniques, vortex generator jets 共VGJs兲 have
frequency range of 0.1⬍ F+ ⬍ 7.7. The dimensionless forcing fre-
shown considerable promise in both steady and unsteady applica-
quency was defined by Bons et al. 关8兴 as the VGJ forcing fre-
tions. As an active system, VGJs offer the benefit of being adapt- quency normalized by the ratio of average freestream velocity
able to different Reynolds number flows 共i.e., flight conditions兲. 共from the jet location to the trailing edge兲 to the suction surface
Experimental results have revealed that steady VGJs offer sub- length 共from the jet location to the trailing edge兲. Bons et al. 关8兴
stantial separation control due to the streamwise vortical struc- further showed that the extent of the control was more profoundly
tures, which pull high momentum fluid from the freestream down impacted by the starting and ending of the jet pulse rather than the
into the separated region, re-energizing the flow 关4,5兴. This con- amount of time the jet remained active.
trol has been shown for a wide range of VGJ blowing ratios 关1兴. Previous work with LPT flow control has been conducted in
Experiments have also shown that pulsed vortex generator jets steady flow cascades without accounting for the unsteady nature
are effective at controlling boundary layer separation for a wide of the flow in an actual engine. In a LP turbine, unsteady distur-
range of operating parameters. The mechanisms of control for bances are continually produced by the upstream blade row. Un-
pulsed VGJs are currently not completely understood. Computa- steady wakes have been shown to re-energize separation regions
tional studies performed by Postl et al. 关6兴 suggested that the as they convect downstream. Stieger et al. 关10兴 attributed this
primary mechanism of control for unsteady VGJs was due to the effect to boundary layer embedded vortical structures. They first
boundary layer transition rather than streamwise vortical struc- noted large amplitude pressure fluctuations as a result of these
tures. These results were obtained at VGJ blowing ratios below wake-induced vortical structures. Later, these structures were
unity. Postl et al. 关6兴 did note that vortical structures began to play identified using PIV. Stieger et al. 关10兴 hypothesized that these
a more important role as the blowing ratios were increased. They vortical structures were created by a rollup of the separated shear
also noted the formation of a 2D 共spanwise兲 disturbance in the layer induced by the wake disturbance.
separation bubble. This disturbance formed after VGJ actuation Gostelow et al. 关11兴 also observed this effect using wake dis-
and helped to accelerate reattachment. Subsequently, Bloxham et turbed flow over a flat plate with an imposed pressure distribution.
al. 关7兴 performed experiments validating some of these conclu- The pressure distribution was representative of the diffusion dis-
sions. tribution seen on a compressor blade and encouraged the devel-
Bons et al. 关8兴 studied the impact of unsteady VGJs on a sepa- opment of a laminar separation bubble. An upstream rod, parallel
ration bubble using the Pack B blade profile. They used boundary with the leading edge of the flat plate, was fastened to a rotating
layer traverses and static pressure taps to monitor the changes in disk. The disk rotated at a rate of 60 rpm, thereby creating two
the separation zone with both steady and unsteady VGJ controls. different wakes 共one from the rod at an upstream location and the
They reported reductions in the wake loss profile of over 50% second from a downstream location兲 every second. Gostelow et al.
with unsteady control, which was later substantiated by the results 关11兴 collected their data by traversing a single-element hot wire
Volino 关9兴 obtained using synthetic jets. The unsteady result ob- through the separation bubble at discrete locations. They showed
that the wake-induced disturbance stabilized the boundary layer.
The wake-induced disturbance was followed by a calmed region
Contributed by the International Gas Turbine Institute of ASME for publication in that delayed transition and stabilized the boundary layer against
the JOURNAL OF TURBOMACHINERY. Manuscript received August 14, 2007; final manu-
script received August 29, 2007; published online February 3, 2009. Review con-
separation. This result was further substantiated by similar studies
ducted by David Wisler. Paper presented at the ASME Turbo Expo 2007: Land, Sea recently performed by Funazaki et al. 关12兴 and Cattanei et al. 关13兴.
and Air 共GT2007兲, Montreal, Quebec, Canada, May 14–17, 2007. Given these well documented effects of wakes on separated

Journal of Turbomachinery Copyright © 2009 by ASME APRIL 2009, Vol. 131 / 021019-1

Downloaded 08 Mar 2009 to 194.225.236.227. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 2 VGJ exit velocity profile and data acquisition locations
„PIV…. VGJ orientation and coordinate system.

blade at 59% Cx. The jets are injected into the flow at a 30 deg
pitch angle and a 90 deg skew angle to the flow as seen in the
inset of Fig. 2. The pressure cavity is connected to high pressure
Fig. 1 Three blade linear cascade air with an inline solenoid valve that regulates the exit velocity of
the VGJs. The jet blowing ratio in this study was fixed near
Bmax = 2.5, where the blowing ratio is defined as the ratio of the jet
flows, it is clear that any active flow control scheme must be exit velocity to the local freestream velocity 共Ujet / Ue at 59% Cx兲.
compatible with the inherently unsteady flow environment in the The jet profile was measured as the VGJ exited the blade into a
LTP. To date, the synchronization of unsteady wakes and unsteady quiescent environment using a single-element hot-film anemom-
control had not previously been investigated. One of the primary eter positioned normal to the jet exit. The resulting jet history plot
objectives of this study was to identify a blowing ratio, jet dura- is presented in Fig. 2. The jet profile is essentially a step function
tion, and synchronization between the unsteady wake disturbance with the initial, high-frequency oscillations attributed to the com-
and the unsteady jet disturbance that caused the greatest time- pressibility of the air in the pressurized cavity. Also featured in
averaged reduction of the separation bubble. This objective was Fig. 2 are the times of data acquisition for the PIV data.
accomplished using static pressure taps and boundary layer data A wake generator is placed 12.7 cm 共0.53 Cx兲 upstream of the
obtained with a single-element hot film. Upon completion of the cascade inlet. A CAD model of the wake generator and its position
optimization study, phase-locked and time-resolved PIV and hot- in the tunnel can be seen in Fig. 3.
film data were taken to identify the relative impacts of the two Unsteady wake disturbances are created using 6 mm diameter
unsteady disturbances and to identify the flow physics that deter- carbon fiber rods. The rods are oriented in the spanwise direction
mined the optimal conditions. and are drawn through the tunnel on a chain sprocket system
driven by a variable frequency motor. Low density foam is used at
Experimental Configuration both the tip and base of the rods to dampen vibrations and seal the
A detailed description of the cascade facility used for this study tunnel. An optical sensor detects the passage of the rods as they
is found in Eldredge and Bons 关14兴. The open-loop wind tunnel is exit the tunnel 共see note in Fig. 3兲 and sends a signal to the
powered by a centrifugal blower. After passing through a series of Parker–Hannifin pulse driver 共t = 0 in Fig. 2兲. This pulse driver
flow conditioners, the air enters an acrylic duct with a velocity controls a solenoid valve used to actuate the VGJs. The pulse
uniformity of ⫾2%. The duct has a cross-sectional area of driver is used to set the duration of the VGJ pulse and the time of
0.15 m2. A square-bar passive grid is placed 5.2 axial chords up- actuation relative to the input signal from the rod sensor 共t = 0兲.
stream of the test section to produce 3% freestream turbulence at The speed of the rods was adjusted to maintain a normalized
the cascade inlet. velocity near Urod / Uin = 0.95 共flow coefficient, ␸ = 0.85兲 with a
The test section is a two passage cascade containing the Pratt & fluctuation of approximately ⫾2%. The period between rods was
Whitney Pack B blade configuration. A depiction of the cascade is measured to be 225 ms. Since the VGJs are synchronized to the
found in Fig. 1. The Pack B blade has an axial chord of 0.238 m, rod passing frequency, this wake period yields a dimensionless
a span of 0.38 m, a design Zweifel coefficient of 1.15, and pro-
vides a cascade solidity of 1.14. At Reynolds numbers below
20,000 共based on inlet velocity and axial chord兲, a nonreattaching
separation bubble forms on the aft portion of the blade beginning
near 68% Cx. The innermost blade in the cascade contains 13
static pressure taps. The taps are located near midspan and are
used to provide a C p profile of the suction surface of the blade.
The C p profile is produced by sequentially connecting these pres-
sure taps to a 0.1 in. H2O Druck differential pressure transducer
referenced to a pitot tube located upstream of the cascade inlet.
This differential pressure is then divided by the dynamic pressure
at the inlet to yield C p. The resulting C p distribution was com-
pared to the prediction generated by the Air Force Research Labo-
ratory using a 2D viscous solver 共VBI, Rao et al. 关15兴兲. After the
blades were geometrically positioned, adjustments in the location
of the tailboards and inlet bleeds were made to most nearly ap-
proximate the VBI solution at high 共nonseparating兲 Reynolds
numbers.
The inner blade of the cascade houses a pressure cavity, which Fig. 3 CAD model of wake generator and test section of tun-
connects to a spanwise row of VGJs. These jets are 2.6 mm in nel. Curved white arrows indicate direction of rotation. Straight
diameter 共d兲 and are spaced 10d apart along the full span of the arrow represents location of optical sensor.

021019-2 / Vol. 131, APRIL 2009 Transactions of the ASME

Downloaded 08 Mar 2009 to 194.225.236.227. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 5 The in-plane PIV configuration. The green plane is a
representation of the laser sheet.

rod using the optical sensor. Figure 2 shows the 15 nondimen-


sional time locations of PIV data acquisition. Time “zero” corre-
sponds with the rod passing through the optical sensor located
outside of the cascade flow path. The first data set was collected at
a nondimensional time t / T = 0.044 共where t is the time relative to
the signal from the optical sensor, and T is the wake-passing pe-
riod兲. The subsequent data sets were taken 15 ms apart. For each
time, data were taken in the upstream and downstream windows at
all 18 z locations. At each location, window and nondimensional
time, 40 images were taken, processed, and averaged. It was pre-
viously shown that averaging with more than 40 images made no
notable difference in the average velocity field results 关16兴. Vector
processing was initially performed with 64⫻ 64 pixel interroga-
tion windows. The interrogation windows were then refined to
32⫻ 32 pixels. A 50% overlap was employed during the vector
Fig. 4 „a… The primary PIV configuration depicting both data processing. According to LaVision 关17兴, the uncertainty in the
regions. „b… The coordinate system used to present the data. seed particle displacement is approximately 0.2 pixel. This trans-
Also included are the merged camera view fields, the axial lates to a velocity uncertainty of ⫾0.08 m / s. The resulting 3D
chord lines of the Pack B, and a representation of the separa-
tion bubble.
blocks of data provide u and v velocity data.
It should be noted that this set of PIV data is presented in the
camera coordinate system. In the region of interest 共59–100%
axial chord兲, the blade is relatively flat. The result is that the x and
forcing frequency of F+ = 0.27 for this study. The rods are spaced y coordinates of the camera are approximately streamwise and
at L / S = 1.64, where L is the distance between the rods and S is the surface normal in this region 共though not exactly兲.
blade spacing. The larger spacing between rods 共compared to the A secondary set of PIV data was used to capture the VGJ-
cascade spacing兲 is intended to simulate vane wakes impinging on induced streamwise vorticity in the plane of the laser sheet. In this
a rotor blade row since the vane count is typically 60–75% of the configuration, the laser was placed on a traverse below the test
blade count for a given LPT stage. section, introducing the laser sheet in the y-z plane, as shown in
Data were taken using a LaVision PIV system mounted to a Fig. 5. The laser created 1.5 mm, consecutive laser sheets with a
three axis traverse below the test section. A Nd:YAG 共yttrium time separation of 150 ␮s. A smaller time separation was required
aluminum garnet兲 laser was used to project two consecutive 1 mm to ensure that a majority of the seed remained in the laser sheet. A
thick laser sheets 共with 250 ␮s time separation兲 in the x-y plane single high-speed camera was placed downstream of the cascade
into the test section 共see Fig. 1 for coordinate system兲. The flow outside of the flow path. Data were collected over four VGJ
was seeded with olive oil particles having diameters between pitches 共⬃100 mm兲. Measurements were taken at five x / d loca-
1 ␮m and 2 ␮m. A high-speed digital camera was positioned be- tions 共10, 15, 20, 35, and 43兲 in order to track the vortical struc-
low the test section. The camera has a resolution of 1376 ture from near inception until it interacts fully with the separation
⫻ 1040 pixels. Measurements were taken in 18 spanwise 共z兲 loca- zone. At each x / d position, eight phase-locked data sets were
tions. The z locations were 1.5 mm apart and spanned one VGJ measured ranging from 20 ms to 100 ms. The data sets were
hole pitch. The first z location was taken directly below a midspan phase locked to the VGJ pulse 共B = 2, 25% duty cycle兲.
VGJ where the flow was shown to be spanwise uniform. Subse- At each location and nondimensional time, 100 images were
quent levels were taken by traversing toward the top of the test taken, processed, and averaged. Vector processing was initially
section in the negative z direction according to the right hand rule performed with 64⫻ 64 pixel interrogation windows. The interro-
共x is the flow direction and y is normal to the blade surface兲. This gation windows were further refined to 16⫻ 16 pixels. A 50%
data set required two different test windows to capture flow along overlap was employed during the vector processing.
the entire blade. These windows covered an upstream 共⬃50% to Prior to taking phase-locked PIV data, the jet duration, blowing
⬃81% Cx兲 and a downstream 共⬃80% to ⬃100% Cx兲 portion of ratio, and time delay 共between the optical sensor signal and VGJ
the blade with approximately 6 mm of overlap 共see Fig. 4共a兲兲. The actuation兲 were optimized to achieve the greatest extent of time-
data windows were later merged together to create a continuous averaged separation bubble reduction. C p distributions were used
set of data as depicted in Fig. 4共b兲. to measure the impact of each of these parameters over a broad
The rod passing period of 225 ms was divided into 15 equal range of values 共time delay of 50– 150 ms, jet duration of
segments. Each segment was phase locked to the passing of the 15– 50 ms, and blowing ratio 1.7–2.5兲. The C p comparisons led to

Journal of Turbomachinery APRIL 2009, Vol. 131 / 021019-3

Downloaded 08 Mar 2009 to 194.225.236.227. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Table 1 Test matrix for synchronization parameter study Table 2 Normalized results from the integrated boundary layer
momentum flux loss parameter
Jet duration Time delay Blowing
Case 共ms兲 共ms兲 ratio Jet Time
duration delay Blowing
No control 共1兲 — — — Case 共ms兲 共ms兲 ratio ⌫ / ⌫o
Wakes only 共2兲 — — —
VGJs only 共3兲 50 N/A 2.5 No control 共1兲 — — — 1
4 50 50 2.5 wakes only 共2兲 — — — 0.75
5 50 100 2.5 VGJs only 共3兲 50 N/A 2.5 0.68
6 50 150 2.5 4 50 50 2.5 0.65
7 50 150 1.7 5 50 100 2.5 0.62
8 30 150 2.5 6 共Optimum兲 50 150 2.5 0.55
7 50 150 1.7 0.63
8 30 150 2.5 0.59

the selection of a smaller range of parameters to be used for quali-


tative comparisons. The test matrix for the qualitative compari-
sons is presented in Table 1. Once the test matrix was selected, the the Pack B blade for each test case.
integrated boundary layer momentum flux losses were calculated The no control C p data lie well below the VBI prediction. The
with 共from Olson et al. 关18兴兲 separation zone is depicted by the relatively flat region in the C p


␦ distribution from 70% to 90% axial chord. The introduction of
Ue2 − U共y兲2
⌫= dy 共1兲 unsteady VGJ control 共Case 3兲 eliminates a portion of this flat-
0
Uin2 tened region, suggesting reattachment of the separation bubble
This calculation provides an estimate of total pressure loss in the near 80% axial chord. For this case, the VGJs had a blowing ratio
suction surface boundary layer and was used to compare the rela- of Bmax = 2.5, a jet duration of 50 ms, and a duty cycle of 23%
tive momentum flux losses for each of the cases listed in the test 共where duty cycle is the ratio of jet duration to the period兲. The
matrix. Several boundary layer velocity profiles were collected at unsteady wake configuration 共Case 2兲 resembles the VBI more
the same location in the separation region 共⬃87% axial chord兲 than the unsteady jet results in the region from 70% to 80% axial
using a single-element hot-film 共uncertainty in velocity of chord but also has a larger deviation from 80% to 90% axial
⫾0.03 m / s兲. The profiles were taken near midspan four jet diam- chord. The addition of VGJs to the unsteady wakes 共Case 6兲 fur-
eters above the bottom edge of a VGJ. ther enhances the control achieved by the unsteady wakes or jets
exclusively. This enhancement was seen over the entire range of
the measured separation zone from 70% to 90% axial chord. As
Results mentioned earlier, the parameters used for the combined unsteady
VGJ Optimization With Wakes. Time-averaged C p distribu- wake and jet C p distribution in Case 6 共B = 2.5, jet duration
tions 共uncertainty in C p of ⫾0.12兲 are presented in Fig. 6 for four = 50 ms, and time delay= 150 ms兲 were determined following a
of the eight test cases in Table 1 共not shown are Cases 4, 5, 7, and rigorous optimization study.
8兲. The solid lines representing the VBI and MISES predictions The C p distribution results suggest that synchronization of un-
are also included. In this C p comparison, the VBI is used as the steady wakes and VGJs is beneficial but does not give any indi-
benchmark of nonseparated flow over the turbine, since it is for a cation as to how sensitive these optimal conditions are to varia-
high, nonseparating Reynolds number. The MISES prediction is tions in the control variables. The integrated boundary layer
included because it is a better representation of the expected C p momentum flux loss parameter 共⌫兲 was used to quantify the con-
distribution at lower Re numbers without control 共no jets or trol effectiveness. The normalized results are tabulated below in
wakes兲. C p distributions that closely resemble the VBI are consid- Table 2. A comparison of the normalized boundary layer momen-
ered to be attached, while deviations from the VBI are indicative tum flux loss parameters for wakes only and VGJs only 共⌫ / ⌫o
of boundary layer separation. The symbols represent the C p from = 0.75 versus 0.68, respectively兲 shows that unsteady VGJs have a
each static pressure tap along the suction and pressure surfaces of more pronounced impact on the momentum flux losses 共separation
region兲. This was an unexpected result given that the unsteady
wake disturbance is a spanwise event while the VGJ disturbance
is not. However, since the z / d location where the boundary layer
共and thus ⌫兲 data were taken aligned directly with the VGJ trajec-
tory 共z / d = 6兲, it is expected that the relative advantage of the VGJ
only case would decrease if the same measurements were taken at
other z / d locations less influenced by the jet. This is due to the
three-dimensionality of the VGJ disturbance and its effect on the
separation bubble dynamics, as will be shown later.
A number of other important synchronization factors can be
gleaned from this study. Three time delays were tested while hold-
ing the jet duration and blowing ratio constant. It is evident that
the largest time delay 共150 ms兲 resulted in the greatest momentum
flux loss reduction 共Case 6 共⌫ / ⌫o = 0.55兲 compared to Case 5
共⌫ / ⌫o = 0.62兲 and Case 4 共⌫ / ⌫o = 0.65兲兲. This would suggest that
the timing between the passing wake and the VGJ disturbances is
an important factor in identifying an optimal synchronization con-
dition. Once the “optimal” time delay was determined, a study
was performed to identify the separation bubbles’ dependence on
Fig. 6 Experimental Cp distributions for the Pack B compared the jet duration. Jet durations of 50 ms and 30 ms were compared
to the VBI. Plot includes no control „no wakes or jets…, wake and resulted in the flux losses, of 0.55 and 0.59, respectively. Jet
only, VGJ only, and combined wakes/jets data. durations larger than 50 ms were not studied to maintain low mass

021019-4 / Vol. 131, APRIL 2009 Transactions of the ASME

Downloaded 08 Mar 2009 to 194.225.236.227. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
flow requirements. These results suggest that jet duration also has the separation bubble is re-energized and pushed off the turbine.
an impact on the flux losses, which corroborates results obtained The vortex cores also migrate away from the jet location in the
by Bloxham et al. 关7兴 for a different blade profile. The final pa- spanwise direction. This movement was expected since the VGJ is
rameter that was adjusted was the blowing ratio. A blowing ratio injected with spanwise momentum. By x / d = 35, the positive vor-
of 2.5 was shown to be significantly more effective at reducing the tex core is positioned near z / d = 5. Given that vortical structures
total pressure losses due to the separation bubble. Higher blowing promote mixing, it should be expected that the separation bubble
ratios were not studied because the maximum allowable pressure would react to the presence of the vortex. Close inspection of the
on the inline solenoid valve was near Bmax = 2.5. three-dimensional nature of the VGJ’s impact on the upstream end
of the separation bubbles 共t / T = 0.84, 0.92, and 0.98 of Fig. 7兲
Detailed Phase-Locked Flow Measurements. Once an opti-
mal synchronizing configuration was obtained, PIV measurements shows that reattachment begins near z / d = 5 and then propagates
were taken to identify the flow physics responsible for the reduced outward. The downwash of the vortex causes the depression in the
momentum flux losses. Comparisons of the wake and jet distur- separation bubble as high momentum fluid is carried into the low
bances were also made. Figure 7 contains isosurfaces of the ve- momentum bubble. Similar VGJ-induced boundary layer modifi-
locity magnitude computed using the PIV data. An isovelocity cations have been observed by Hansen and Bons 关5兴 and Khan
surface of U / Uin = 1.0 was selected because it clearly depicts the and Johnston 关20兴. Although the in-plane PIV data were collected
distinct influences of the passing wake and jet. Each of the 15 data without the addition of passing wakes, similar three-dimensional
acquisition times is represented in the figure, depicting the sepa- structures were seen in both sets of data. These data suggest that
ration bubbles’ behavior over the complete period. Figure 7 also streamwise vortices also participate in the removal of the separa-
includes an isovelocity surface without wakes or VGJs for com- tion bubble.
parison. These surfaces give an indication of the jet and wake Once the separation bubble is ejected from the blade 共t / T
effects on the flow. In order to facilitate identification of the sepa- = 0.24 in Fig. 7兲, there is a period of time before the bubble begins
ration bubble, the curvature of the turbine blade was removed to recover. The isovelocity surfaces at t / T of 0.31, 0.38, and 0.44
from the isovelocity surface height. Thus, the vertical axis 共y / d兲 show very little growth in the separation region. By t / T = 0.51, the
represents the distance from the isovelocity contour to the blade boundary layer begins to separate again. To better understand this
surface. Accordingly, elevated portions of the isovelocity surface period of sustained control, single-element hot-film data were
are attributed to the separation bubble. The flow moves from right taken using a blade follower device. This device keeps the hot
to left as x / d extends from 0 to 67 共approximately 59–100% axial film at a predetermined distance from the wall. The follower is
fixed to a single axis traverse, which allows the hot film to
chord兲. The VGJ is located near a z / d of 9 共hole center兲 but is
traverse from 48% axial chord to the trailing edge. Thirteen pro-
only active in the range of t / T = 0.71– 0.84.
files were taken ranging from 2 mm to 16 mm from the wall.
Since t / T = 0 is referenced to the passing of the rod through the
Each profile was taken at a z / d = 6 共four jet hole diameters above
optical sensor, indication of a passing wake is not immediately
the bottom edge of a VGJ hole兲. Phase-locked data were taken for
evident in the isovelocity surfaces. At t / T = 0.04, the lingering
24 s at 10 kHz 共approximately 106 wake-passing cycles兲. The
effects of a VGJ/separation bubble interaction are still present.
data analysis technique described in Bons et al. 关21兴 was em-
This VGJ pulse started about ⌬t / T = 0.33 共75 ms兲 prior to the
ployed to obtain Urms after removing the phase-locked mean ve-
passing of the rod through the optical sensor 共it is therefore phase
locity from the raw velocity signal.
locked to the previous rod passing兲. The VGJ has caused the sepa-
Urms / Uin data were presented in Fig. 9 to help identify the wake
ration bubble to reattach at the upstream end. The higher momen-
disturbance. Figure 9 is divided into 24 plots representing 24
tum fluid in the reattached region meets the slow moving separa-
phase-locked data windows taken over the wake-passing period
tion bubble and causes an elevated bulge in the isovelocity
共T兲. Similar to the PIV data, the hot-film data were phase locked
surface. This bulge convects off the end of the blade in subsequent
data sets. The three-dimensional effect of the VGJ on the separa- using the rod optical sensor. The nondimensional time is shown in
tion bubble is still very apparent as the separation bubble moves the upper right corner of each plot. The use of Urms / Uin plots
off the blade in t / T of 0.11, 0.18, and 0.24. assists in the identification of the separated flow region, the pulsed
The three-dimensional nature of the jet’s unsteady effect on the jet trajectory, and 共to a lesser extent兲 the wake trajectory. From
separation bubble has previously been attributed primarily to a t / T = 0.04– 0.25, the separation bubble 共x / Cx ⬎ 0.8兲 is decreasing
VGJ-induced transition of the boundary layer 关19兴. In-plane PIV in size due to the influence of the previous VGJ disturbance. The
data were used to clarify the role of vortical structures in the VGJ wake disturbance 共shown as a red arrow兲 then enters the measure-
control. In order to isolate the influence of the VGJ, the in-plane ment domain as evidenced by a slight increase in freestream tur-
data were taken without the wake disturbance. Figure 8 contains bulence upstream of the separation bubble. The separation bubble
streamwise vorticity data 共y-z plane兲 collected 5 ms before the is further reduced in size due to the passing of the wake 共t / T
VGJ deactivated. In this study, the blowing ratio of the VGJ was = 0.54– 0.71兲. Once the wake has passed, there remains a region of
Bmax = 2, the jet duration was 50 ms, and the duty cycle was 25%. low turbulence referred to by Gostelow et al. 关11兴 as a “calmed
The streamwise vorticity for four x / d locations 共x / d = 10, 20. 25, zone.” This region of low turbulence is seen at the trailing edge
and 35兲 is provided in the figure to track the vortical development. 共x / Cx ⬎ 0.9兲 from t / T = 0.75 until the influence of the jet distur-
The VGJ location is represented in the figure by the red arrow bance arrives 共green arrow兲.
near z / d = 9 共jet hole center兲. The calmed zone is further evident in the time history plot at
The plot of x / d = 10 depicts the strong positive and negative y / d = 0.80 presented in Fig. 10共a兲. The two-sided red arrow iden-
VGJ-induced vorticity cores. The cores are positioned near z / d tifies the calmed zone that results from the wake disturbance. The
= 7 and y / d = 2. These strong vorticity cores dissipate in the sub- smaller black arrow identifies the calmed zone that results from
sequent plots. Despite the energy dissipation, the positive vortex the VGJ disturbance. Figure 10共b兲 is the time history plot for the
maintains its structure up to x / d = 35 共well into the separation wakes only case. In the absence of an intermediate jet disturbance,
region兲. As the vortical structures move downstream, they migrate this plot shows bubble regrowth 共Urms / Uin ⬎ 5 % 兲 beginning at
away from the wall. By x / d = 35, the positive core has migrated t / T = 1.1. It appears that the VGJ disturbance arrives at the sepa-
out to y / d = 4. Close examination of the isovelocity surfaces pre- ration bubble just prior to the breakdown of the calmed zone
sented previously in Fig. 7 共t / T = 0.84兲 suggests that the vortex is caused by the wake. This new disturbance prevents regrowth of
migrating away from the wall due to the presence of the separa- the separation bubble and produces another calm zone. A short
tion bubble. The in-plane PIV data for subsequent time steps 共not time later, a new wake disturbance re-establishes the wake-
presented兲 show the vortex core migrate back toward the wall as induced calmed zone and the cycle continues. These figures sug-

Journal of Turbomachinery APRIL 2009, Vol. 131 / 021019-5

Downloaded 08 Mar 2009 to 194.225.236.227. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 7 Phase-locked isovelocity surfaces „U / Uin = 1.0… for wakes/jets „Case 6… configuration. The red arrows indicate ap-
proximate jet locations.

gest that the optimal synchronization of jets/wakes prolongs the induced calm zone.
calm zone and suppresses separation bubble regrowth. In order to Returning to Fig. 7, the separation bubble begins to recover
optimize the control of wakes/jets, the jet disturbance should in- after the jet-induced calm zone. Then, between t / T = 0.51 and
teract with the separation zone just prior to the end of the wake- 0.58, the separation bubble is impacted by the wake disturbance

021019-6 / Vol. 131, APRIL 2009 Transactions of the ASME

Downloaded 08 Mar 2009 to 194.225.236.227. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Fig. 8 Streamwise vorticity comparison for VGJs only „Case
3…. VGJ at x / d = 0 and z / d = 9 „hole center…. Blowing ratio, Bmax
= 2.

共see Fig. 9兲. The wake interaction with the separated zone results
in a spanwise-uniform bulge in the isovelocity surface that is sub-
sequently carried off the blade by t / T = 0.91. The upstream end of Fig. 10 Time history plots „Urms / Uin… depicting wake/jet and
the residual separation bubble is then impacted by the next VGJ wake only interaction with the separation bubble
disturbance.
After the wake passes in the plot of t / T = 0.91, there is a sig-
nificantly larger separation bubble in comparison to the residual grated and normalized by the no control case. Figure 11 is a plot
bubble after the jet disturbance 共t / T = 0.31– 0.51兲. In order to of this integrated measurement for each of the nondimensional
quantify the size of the separation bubbles at each t / T, each is- times 共wakes/jet and wakes only data兲.
ovelocity surface from Fig. 7 was averaged in the spanwise direc- Figure 11 shows the impact of each of the disturbances and
tion. The resultant average isovelocity surfaces were then inte- their relative effectiveness in suppressing the separation bubble.

Fig. 9 Urms / Uin plots of the wakes/jets „Case 6… configuration. The nondimensional time
is labeled in the upper right corner of each plot.

Journal of Turbomachinery APRIL 2009, Vol. 131 / 021019-7

Downloaded 08 Mar 2009 to 194.225.236.227. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Cx ⫽ axial chord 共24 cm兲
C p ⫽ pressure coefficient 共Ptin-P兲 / 共Ptin-Psin兲
F+ ⫽ dimensionless forcing frequency
共f / 共Uav / SSLJ兲兲
L ⫽ distance between rods
P ⫽ pressure
Rec ⫽ Reynolds number based on inlet velocity and
axial chord 共CxUin / ␯兲
S ⫽ blade spacing
SSLJ ⫽ suction surface length from jet location to the
trailing edge
Fig. 11 Integrated isovelocity surfaces „Cases 6 and 2… at each U ⫽ velocity magnitude
data acquisition time. The data were normalized by the size of Uav ⫽ average freestream velocity from jet location to
the no control separation bubble. the trailing edge
T ⫽ rod passing period 共225 ms兲
d ⫽ jet hole diameter 共2.6 mm兲
The configuration with wakes only causes a decrease in the nor-
malized separation zone from 0.94 to 0.72. At t / T = 0.78, the nor- f ⫽ VGJ forcing frequency
malized separation bubble grows to nearly 0.81 as the 2D wake t ⫽ time 共s兲
disturbance impacts it. The bubble size then decreases to 0.58 as u ⫽ x-component of velocity 共approximately
the 2D disturbance is ejected from the blade. The average size of streamwise兲
the separation bubble decreases very rapidly as evidenced by the v ⫽ y-component of velocity 共approximately blade
slope of the line during wake-induced control. A slower reduction normal兲
is noted in the VGJ-induced control. A comparison of the speed w ⫽ z-component of velocity 共spanwise兲
and size of these reductions indicates that the spanwise-average x ⫽ approximate streamwise coordinate
wake-induced control might actually have more impact than the y ⫽ approximate blade normal coordinate
jet. After the wake passes, the jet disturbance interacts with a z ⫽ spanwise coordinate 共z = 0 at bottom of VGJ
partial separation bubble. The remainder of the low momentum hole兲
fluid is re-energized, further decreasing the separation bubble to ␦ ⫽ boundary layer thickness
0.42 共0.3 less than the wakes only configuration兲. These results ␾ ⫽ flow coefficient 共Uin,axial / Urod兲
suggest that at the optimal synchronizing configuration the wake ⌫ ⫽ integrated boundary layer momentum flux loss
disturbance prepares the separation bubble for maximum jet effec- ␯ ⫽ kinematic viscosity
tiveness.
Subscripts
Conclusions axial ⫽ axial direction
e ⫽ boundary layer edge
Surface static pressure and hot-film data were used to identify in ⫽ cascade inlet conditions
“optimal” conditions for the synchronization of VGJ and wake jet ⫽ VGJ jet
disturbances. Results suggest that jet duration, blowing ratio, and max ⫽ max
the time delay between disturbances all have a significant impact o ⫽ base line case without wakes or VGJ control
on control effectiveness. Single camera PIV and hot-film data rod ⫽ wake generator rod
were used to identify the relative impacts of the two unsteady rms ⫽ root mean square
disturbances and the flow physics that resulted in the control ef- S ⫽ static
fectiveness. In-plane PIV data showed that the three-dimensional T ⫽ total
shape of the jet-disturbed separation bubble coincided with the
location of a streamwise vortical structure. The depression in the
separation bubble corresponded with the downwash of the vortical References
structure. Hot-film and PIV data were used to show that the wake
关1兴 Sondergaard, R., Bons, J. P., and Rivir, R. B., 2002, “Control of Low-Pressure
and jet disturbances produced calmed zones. At optimal condi- Turbine Separation Using Vortex Generator Jets,” J. Propul. Power, 18共4兲, pp.
tions, the jet disturbance arrived at the separation bubble just prior 889–895.
to the breakdown of the wake-induced calmed zone. Conse- 关2兴 Sharma, O., 1998, “Impact of Reynolds Number of LP Turbine Performance,”
quently, the jet disturbance interacted with a smaller separation Proceedings of 1997 Minnowbrook II Workshop on Boundary Layer Transition
I Turbomachines, NASA/CP-1998-206958, pp. 65–69.
bubble. This resulted in the most substantial removal of the sepa- 关3兴 Matsunuma, T., Abe, H., Tsutsui, Y., and Murata, K., 1998, “Characteristics of
ration zone. The location of the VGJs in this study was based on an Annular Turbine Cascade at Low Reynolds Numbers,” ASME Paper No.
separation predictions in a wake-free environment. The addition 98-GT-518.
of the unsteady wakes moved the separation bubble further down- 关4兴 Eldredge, R. G., and Bons, J. P., 2004, “Active Control of a Separating Bound-
ary Layer With Steady Vortex Generating Jets—Detailed Flow Measure-
stream away from the VGJs. The vortex generator jets would ments,” presented at the AIAA Aerospace Sciences Meeting, Reno, NV, Jan.
likely be more effective, and consequently require less mass flow, 5–8.
if placed closer to the new separation region. 关5兴 Hansen, L. C., and Bons, J. P., 2006, “Phase-Locked Flow Measurements of
Pulsed Vortex-Generator Jets in a Separating Boundary Layer,” J. Propul.
Power, 22共3兲, pp. 558–566.
Acknowledgment 关6兴 Postl, D., Gross, A., and Fasel, H. F., 2003, “Numerical Investigation of Low-
Pressure Turbine Blade Separation Control,” AIAA Paper No. 2003-0614.
This research could not have been performed without the spon- 关7兴 Bloxham, M. Reimann, D., and Bons, J. P., 2006, “The Effect of VGJ Pulsing
sorship of the Air Force Office of Scientific Research, with Dr. Frequency on Separation Bubble Dynamics,” presented at the AIAA 44th
Thomas Beutner and Lt. Col. Rhett Jefferies as program manag- Aerospace Sciences Meeting and Exhibit, Reno, NV, Jan. 9–12, Paper No.
ers. AIAA 2006-0876.
关8兴 Bons, J. P., Sondergaard, R., and Rivir, R. B., 2002, “The Fluid Dynamics of
LPT Blade Separation Control Using Pulsed Jets,” ASME J. Turbomach., 124,
Nomenclature pp. 77–85.
关9兴 Volino, R. J., 2003, “Separation Control on Low-Pressure Turbine Airfoils
B ⫽ blowing ratio 共Ujet / Ue at 59% Cx兲 共use Umax if Using Synthetic Vortex Generator Jets,” ASME Paper No. GT2003-38729.
pulsed VGJs兲 关10兴 Stieger, R., Hollis, D., and Hodson, H., “Unsteady Surface Pressures Due to

021019-8 / Vol. 131, APRIL 2009 Transactions of the ASME

Downloaded 08 Mar 2009 to 194.225.236.227. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm
Wake Induced Transition in a Laminar Separation Bubble on a LP Turbine 关16兴 Bons, J. P., Hansen, L. C., Clark, J. P., Koch, P. J., and Sondergaard, R., 2005
Cascade,” ASME Paper No. GT2003-38303. “Designing Low-Pressure Turbine Blades With Integrated Flow Control,”
关11兴 Gostelow, J. P., and Thomas, R. L., 2003, “Response of a Laminar Separation ASME Paper No. GT2005-68962.
Bubble to an Impinging Wake,” ASME Paper No. GT2003-38972. 关17兴 LaVision, 2004, DAVIS FLOWMASTER, v. 7.0, Mar., LaVision GmbH, Anna-
关12兴 Funazaki, K., Yamada, K., Ono, K., Segawa, K., Hamazaki, H., Takahashi, A., VandenHoeck-Ring, 19, D-37081 Gottingen.
and Tanimitsu, H., “Experimental and Numerical Investigations of Wake Pass- 关18兴 Olson, D. H., Reimann, D., Bloxham, M., and Bons, J. P., “The Effect of
ing Effects Upon Aerodynamic Performance of a LP Turbine Linear Cascade Elevated Freestream Turbulence on Separation Control With Vortex-
With Variable Solidity,” ASME Paper No. GT2006-90507. Generating Jets,” presented at the AIAA 43rd Aerospace Sciences Meeting and
关13兴 Cattanei, A., Zunino, P., Schroder, T., Stoffel, B., and Matyschok, B., 2006, Exhibit, Reno, NV, Jan. 10–13, Paper No. AIAA 2005-1114.
“Detailed Analysis of Experimental Investigations on Boundary Layer Transi- 关19兴 Reimann, D., Bloxham, M., Crapo, K., Pluim, J., and Bons, J. P., “Influence of
tion in Wake Disturbed Flow,” ASME Paper No. GT2006-90128. Vortex Generator Jet-Induced Transition on Separating Low Pressure Turbine
关14兴 Eldredge, R. G., and Bons, J. P., 2004, “Active Control of a Separating Bound- Boundary Layers,” AIAA Fluid Dynamics Conference, San Francisco, Jun.
ary Layer with Steady Vortex Generating Jets—Detailed Flow Measurements,” 5–8, Paper No. AIAA2006-2852.
presented at the AIAA Aerospace Sciences Meeting, Reno, NV, Jan. 5–8, Pa- 关20兴 Khan, Z. U., and Johnston, J. P., 2000, “On Vortex Generating Jets,” Int. J.
per No. AIAA-2004-0751. Heat Fluid Flow, 21, pp. 506–511.
关15兴 Rao, K. V., Delaney, R. A., and Topp, D. A., “Turbine Vane-Blade Interaction: 关21兴 Bons, J. P., Bloxham, M., and Reimann, D., 2006, “Separated Flow Transition
Vol. 1.2-D Euler/Navier-Stokes Aerodynamic and Grid Generation Develop- on an LP Turbine Blade With Pulsed Flow Control,” ASME Paper No.
ments,” U.S. Air Force Research Laboratory, Report No. WL-TR-94-2073. GT2006-90754.

Journal of Turbomachinery APRIL 2009, Vol. 131 / 021019-9

Downloaded 08 Mar 2009 to 194.225.236.227. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

You might also like